Вы находитесь на странице: 1из 12

Applied Catalysis A: General 288 (2005) 220–231

www.elsevier.com/locate/apcata

Mg–Fe–Al mixed oxides with mesoporous properties prepared


from hydrotalcite as precursors: Catalytic behavior
in ethylbenzene dehydrogenation
Yoshihiko Ohishi a, Tomonori Kawabata a,
Tetsuya Shishido b, Ken Takaki a, Qinghong Zhang c,
Ye Wang c, Kiyoshi Nomura d, Katsuomi Takehira a,*
a
Department of Chemistry and Chemical Engineering, Graduate School of Engineering, Hiroshima University,
Kagamiyama 1-4-1, Higashi-Hiroshima 739-8527, Japan
b
Department of Chemistry, Tokyo Gakugei University, Nukui-kita 4-1-1, Koganei, Tokyo 184-8501, Japan
c
State Key Laboratory for Physical Chemistry of Solid Surfaces, Department of Chemistry, Xiamen University, Xiamen 361005, China
d
School of Engineering, University of Tokyo, Hongo 7-3-1, Bunkyoku, Tokyo 113-8656, Japan
Received 18 January 2005; received in revised form 15 April 2005; accepted 27 April 2005
Available online 6 June 2005

Abstract

Supported iron catalyst was prepared from Mg–Fe–Al hydrotalcite-like compounds as precursors and successfully applied for the
ethylbenzene dehydrogenation to styrene. After the calcination, the iron-substituted hydrotalcite-like samples were converted to mixed oxides
with a high surface area as well as a mesoporous character; the XRD analysis indicates the formation of periclase Mg(Fe, Al)O as a main
phase. Catalytic tests of the Mg2FexAl1x catalysts showed that the styrene conversion increased with increasing the iron content up to
x = 0.75 and then decreased, while the selectivity was the highest at x = 0.25. The optimum temperature for the reaction was 550 8C, which
was lower than that used in the commercial process. No favorable effect of the addition of either CO2 or O2 in the reaction medium was
observed in this reaction. Actually, the pre-treatment with H2 resulted in an increase in the activity at the beginning of the reaction as well as a
stable activity during the reaction. The ethylbenzene conversion of 60% and the styrene selectivity of 95% were kept for 3 h over
Mg3Fe0.5Al0.5 catalyst at 550 8C. After the reaction for 3 h, the iron species on the catalyst was partially reduced to the valence state between
Fe2+ and Fe3+. The high catalytic performance can probably be attributed to the formation of partially reduced iron oxides on the surface of
catalyst and to the high surface area along with the porous structure, which originated from the Mg–Fe–Al hydrotalcite structure in the
precursors.
# 2005 Elsevier B.V. All rights reserved.

Keywords: Mg–Fe–Al hydrotalcite-like materials; Ethylbenzene dehydrogenation; The reduced iron oxide; Porous catalyst

1. Introduction where the thermal cracking becomes significant [1,2]. The


main precursors of the technical catalyst are hematite
Styrene, an important basic chemical as a raw material for (Fe2O3) and the promoter potassium carbonate (K2CO3);
polymers, is produced commercially by the dehydrogena- these are mixed and calcined. Small amounts of other metal
tion of ethylbenzene using a typical Fe-K-Cr oxide-based oxides like Cr2O3 are added as structural promoters to
catalyst in the presence of a large quantity of steam at high stabilize the catalyst morphology and to prevent sintering.
temperatures of 600–700 8C, just below the temperature Promotion of iron oxide with potassium enhances the
reactivity of iron oxide by an order of magnitude and reduces
* Corresponding author. Tel.: +81 824 24 7744; fax: +81 824 24 7744. the formation of carbonaceous surface deposits or coke that
E-mail address: takehira@hiroshima-u.ac.jp (K. Takehira). deactivates the catalysts [1,3,4]. A study on the catalyst

0926-860X/$ – see front matter # 2005 Elsevier B.V. All rights reserved.
doi:10.1016/j.apcata.2005.04.033
Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231 221

structure and composition under reaction conditions by compounds as precursors for mixed oxide catalysts in
Muhler et al. [5,6] gave evidence for the existence of KFeO2 various reactions as reviewed by Cavani et al. [22]. Indeed,
which, in line with a proposal of Hirano [4], was interpreted Clause et al. [23] reported that the catalysts obtained by
as essential for high activity. KFeO2 was identified as the thermal decomposition of Ni–Al hydrotalcite precursors
active phase also by Lundin et al. [7] and Coulter et al. [8]. have very interesting properties, such as high thermal
Kinetic studies on model catalysts showed similar apparent stability and good metal dispersion. The latter behavior has
activation energies for K-promoted and unpromoted iron been attributed by Kruissink et al. [24] to the particular
oxide catalysts, suggesting that potassium only increases the structure of the hydrotalcite through a homogeneous
number of active sites, but does not change their nature [8]. distribution of the metallic cations in the brucite-type
Potassium stabilizes the catalyst against reduction [9] and sheets. The authors have reported that Mg–Ni–Al hydro-
supports the removal of coke. K2CO3 is believed to represent talcite-like compounds were effective as precursors of the
the active center for carbon gasification [4,10,11], whereas catalyst for steam and autothermal reforming of CH4
benzene and toluene as the main side products may be [25,26]. A new catalyst for the process of ethylbenzene
formed at different sites [4,8,12]. Experiments on unpro- dehydrogenation to styrene under a CO2 or O2 atmosphere
moted Fe2O3 gave clear evidence that defective surfaces are was obtained using vanadium-substituted Mg–Al hydro-
more active than well-ordered ones [13]. Ranke and co- talcite-like compounds as precursors [27,28]. Fe–Mg mixed
workers [14] reported that both unpromoted Fe2O3 and K- oxides with high surface areas around 200 m2 g1 and with
promoted model catalysts show a similar high starting mesopores around 15 nm in diameter were prepared from
activity, while that of Fe3O4 is clearly lower. Catalyst hydrotalcite-like compounds and tested as the catalysts in
deactivation by coking or oxide reduction was prevented by the Fisher–Tropsch reaction [29]. Moreover, surface acid/
the addition of potassium to the catalyst, and by the addition base properties and the reducibility of the iron of the
of water or a small amount of oxygen to the feed. catalysts were studied by microcalorimetric method and
Although the commercial catalysts are very active and Mössbauer spectroscopy [30]. Fe–Mg or Fe–Mg–Al mixed
selective, they have some disadvantages: (i) The active oxides were prepared from hydrotalcite-like precursors and
oxidation state is unstable; hematite (a-Fe2O3) is preferred were successfully tested as the catalysts for the reduction of
for styrene production, but tends to lower oxides and even to aromatic nitro compounds [31–33]. As for the dehydrogena-
elemental iron, both of which catalyze coke formation and tion of ethylbenzene, Fe–Mg–Al [34], Fe–Mg–Zn–Al [35],
dealkylation [15,16]. (ii) The catalysts have low surface and Fe–Zn–Al [36] mixed oxides were tested under a CO2
area. (iii) They are deactivated with time, being susceptible atmosphere. The activity losses over the Fe-Mg-Al catalysts
to poisoning by halides and residual organic chloride with time-on-stream were completely restored by oxygen
impurities [1]. The most serious deactivation is caused by pulses [34]. The Fe-Mg-Zn-Al catalysts afforded the highest
the loss of potassium promoter, which migrates in two ehylbenzene conversion of 53.8% and a styrene selectivity
directions as the catalyst ages. Potassium chlorides are found of 96.7% at 500 8C [35]. The Fe-Zn-Al catalysts were
down-stream in the water layer of the condensed product as effective and gave an areal rate of 4.15 mmol min1 m2
well as in catalyst pellets. In fact, a major migration of none the less the surface area was as low as 22.1 m2 g1
cat [36].
potassium occurs within the catalysts pellets, because the In the present work, a series of Fe–Mg–Al mixed oxides
center of the pellets operates at a slightly lower temperature were prepared from the corresponding hydrotalcite-like
than the periphery due to the endothermicity of the reaction. precursors and were tested for the dehydrogenation of
In addition, potassium migrates also to the cooler part of the ethylbenzene. Catalytic behaviors of Fe–Mg–Al mixed
reactor (exit) [1,15,16]. Also, the toxicity of the chromium oxides are completely different from those reported in the
compounds causes damage to humans and to the environ- preceding papers [34–36]; no favorable effect of either the
ment [1,16]. Therefore, the search for new catalyst systems additions of CO2 and O2 or the oxidation pre-treatment of
which have high surface areas and can stabilize the active the catalyst on the activity was observed, whereas the
state of iron, in the absence of potassium, is much needed. reduction pre-treatment afforded the catalyst with somewhat
Aluminum was proved to be an excellent promoter, higher activity as well as higher stability. The results of
preventing sintering in iron-oxide catalysts [17,18]. Hirano detailed studies on the catalyst structure and the activities
[19] studied the effect of adding a series of alkaline earth will be reported in this paper.
oxides such as MgO, CaO, SrO, and BaO to a K-promoted
iron oxide catalyst, among which especially MgO had good
characteristics. Stobbe et al. [20] have developed methods to 2. Experimental
support promoted iron oxide onto an MgO support material.
A supported system can prevent the physical degradation of 2.1. Catalyst preparation
the catalyst particles.
The starting materials can also play a significant role in The Mg–Fe and Mg–Fe–Al hydrotalcite-like precursors
the catalytic performance [21]. There has been for the past were synthesized using a co-precipitation of nitrates of each
years an increased interest in using hydrotalcite-like metal component, following the method by Miyata and
222 Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231

Okada [37] after minor modifications. An aqueous solution rate of 10 8C min1 from ambient temperature to 1100 8C
containing the nitrates of Mg2+, Fe3+ and Al3+ (300 ml) was using a mixture of 5 vol.% H2/Ar at a rate of 100 ml min1 as
added slowly with vigorous stirring into an aqueous solution reducing gas, after passing through a 13 molecular sieve trap
of sodium carbonate (300 ml). By adjusting the pH of the to remove water. We used a U-shaped quartz tube reactor, the
solution at 10 with an aqueous solution of sodium hydroxide, inner diameter of which was 6 mm; it was equipped with a
we caused a heavy slurry to precipitate. After the solution TCD for monitoring the H2 consumption. Prior to the TPR
was aged at 60 8C for 24 h, the precipitates were washed measurement, 50 mg of sample was calcined at 550 8C for 1 h
with de-ionized water (1000 ml), dried in air at 110 8C for in Ar gas flow at a rate of 20 ml min1. N2 adsorption (77 K)
24 h, and calcined at 550 8C for 12 h in a muffle furnace in a study was used to examine both BET surface area and the
static air atmosphere. As a control, imp-Fe/MgO and imp-Fe/ porous property of the Mg–Fe–Al mixed oxide. The
g-Al2O3 catalysts were prepared by incipient wetness measurement was carried out on a Bell Japan Belsorp
method using magnesia (UBE) and alumina (JRC-ALO8) 18SP instrument (volumetric); all samples were pre-treated in
as the carrier. vacuum at 200 8C for 12 h before the measurements. The
pore-size distribution was evaluated from the adsorption
2.2. Characterizations of catalysts isotherm by the Dollimore and Heal (DH) method [39].

Inductively coupled plasma (ICP) optical emission 2.3. Catalytic reaction


spectroscopy was used for the determination of the metal
content in each sample synthesized above. The measurements Dehydrogenation of ethylbenzene was conducted using a
were performed with a Perkin-Elmer OPTIMA 3000; each fixed-bed flow reactor at atmospheric pressure. A quartz
sample was completely dissolved in a mixture of HF and glass tube with an inner diameter of 8 mm was used as a
HNO3 before the measurements. Powder X-ray diffraction reactor. In the dehydrogenation reactions, typically 0.15 g of
was recorded on a Rigaku powder diffraction unit, RINT catalyst, which had been pelletized and crushed to the
2250VHF, with mono-chromatized Cu Ka radiation particles 250–417 mm in diameter, was loaded into the
(l = 0.154 nm) at 40 kV and 300 mA. The diffraction pattern reactor. The material was treated in a gas flow containing N2/
was identified by comparing with those included in the JCPDS O2 (10/10 ml min1) at 550 8C for 1 h and then purged with
database (Joint Committee of Powder Diffraction Standards). N2. The reaction was started by introducing a gas mixture of
Diffuse reflectance UV–vis spectroscopic measurements ethylbenzene, N2 and CO2, O2 or Ar to the reactor.
were recorded on a JASCO UV/VIS/NIR (V-570) spectro- Ethylbenzene (0.6 ml min1; ca. 1.5 mmol h1) was fed by
photometer. The spectra were collected in the range of 200– bubbling a gas mixture of N2 (10 ml min1) and CO2, O2 or
700 nm referenced to BaSO4. The sample was loaded in an in Ar (30 ml min1) through liquid ethylbenzene held at
situ cell and was treated in a pure N2 gas flow at 200 8C for 28.3 8C in a thermostat. The reaction products (styrene,
dehydration before the measurement. X-ray absorption toluene, and benzene) and ethylbenzene were liquefied in a
spectroscopic measurements were performed at room cold trap of n-heptane and acetone at 0 8C, and analyzed
temperature in a transmission mode at the EXAFS facilities with an FID gas chromatograph using a packed ID-BPX 5
installed at the BL01B1 line of Spring-8 JASRI, Harima, column and cyclohexanone as an internal standard. Analyses
Japan, using a Si(1 1 1) monochrometer. The data were of gaseous products (CO, CO2, and H2) were performed with
collected in a quick-XAFS mode. Data reductions were a TCD gas chromatograph using a packed Molecular Sieve-
performed with the FACOM M1800 computer system of the 5A column and Porapak Q. All the lines and valves between
Data Processing Center of Kyoto University [38]. The sample the cold trap and the reactor were heated to 150 8C to prevent
was mixed with boron nitride as a binder and then pressed into any condensation of ethylbenzene or of the dehydrogenation
a disk (10 mm in diameter). Energy was calibrated with Cu K- products. The presence of other hydrocarbons and of the
edge absorption (8981.0 eV); the energy step for measure- oxygenated products was checked by GC with FID using a
ment in the XANES region was 0.3 V. The adsorption was packed FFAP column, but none were detected. All data of
normalized to 1.0 at an energy position of 30 eV higher than the reactions were collected after the reaction for 1 h and
the adsorption edge. Transmission Mössbauer spectra of used for the calculation of the reaction results unless
pelletized powder samples were recorded at room tempera- specifically mentioned, since the activity of the catalyst
ture, using a constant acceleration mode (Topologic System sometimes gradually decreased during the reaction.
Co.) of a radiation source with about 40 MBq 57Co(Cr) and a
YAP scintillation counter. Doppler velocity was calibrated
with reference to a-Fe. Thermogravimetric-differential 3. Results and discussion
thermal analysis (TG-DTA) of the catalyst was performed
under an inert atmosphere of N2 (20 ml min1) with 3.1. Structure and properties of supported Fe catalysts
Shimadzu TGA-50 and DTA-50 analyzers using 50 mg of
sample at a rate of 10 8C min1. Temperature programmed Metal compositions, surface areas and Fe contents of the
reduction (TPR) of the catalyst was performed at a heating supported Fe catalysts are shown in Table 1. Metal
Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231 223

Table 1 together with a weak line of MgFe2O4 spinel, suggesting that


Surface areas of supported Fe catalysts a reaction between iron and magnesia occurred during
Catalyst Preparation Fe wt.%b Surface area Color impregnation, though slowly. imp-Fe/g-Al2O3 showed
methoda (m2 g1
cat )
c
reflection lines of a-Fe2O3 (hematite) and g-Al2O3 alone,
Mg3Fe0.5Al0.5 cp 21.4 169.0 Off-white and no other line was observed, indicating that no substantial
imp-Fe/MgO imp 21.4 67.9 Off-white reaction between iron and alumina occurred.
imp-Fe/g-Al2O3 imp 21.4 137.1 Reddish brown
Mg2Fe cp 49.8 115.5 Brown
To understand the decomposition procedure of the
Mg2Fe0.75Al0.25 cp 39.1 142.2 Pale brown Mg3Fe0.5Al0.5 HT, we analyzed the TG-DTA profile in
Mg2Fe0.5Al0.5 cp 27.3 170.3 Pale brown detail. As reported in references [22,29], two weight loss
Mg2Fe0.25Al0.75 cp 14.4 174.1 Off-white stages were observed on the TG curve; these correspond to
Mg2Al cp 0 202.9 White the two endothermic peaks around 177 and 366 8C on the
a
cp: Co-precipitation method; imp: impregnation method. DTA profile, demonstrating that the decomposition pro-
b
Measured by ICP method.
c ceeded in two steps. When 10 mg of the sample was used,
Measured by BET method.
the total weight loss was found to be 4.25 mg; 1.53 mg was
lost in the first step, while 2.72 mg was lost in the second
compositions shown as suffixes in each catalyst are calculated step. The sample loses water molecules contained in the
based on the amounts of metal nitrates used in the preparation. interlayer at the first step and hydroxide and carbonate
Higher surface area was observed for the catalysts prepared groups at the second step. Suppose that the initial sample has
from hydrotalcite-like compounds by co-precipitation com- the composition of Mg6Fe1Al1(OH)16CO3xH2O and that
pared with those prepared by impregnation. the final mixed oxide has the composition of 6MgO 1/
XRD patterns of Mg3Fe0.5Al0.5 hydrotalcite (HT) 2Fe2O3 1/2Al2O3, the formula weight of the initial sample
obtained by co-precipitation showed typical reflection lines can be calculated to be 648.1 and the number of water
of layered double hydroxides. After the calcination of molecules contained in the interlayer was found to be 4.9,
Mg3Fe0.5Al0.5 HT at 550 8C in N2 atmosphere (Fig. 1a), the which is approximately equal to 5. Thus, the formula of the
lines of periclase Mg(Fe, Al)O, i.e., periclase MgO phase Mg3Fe0.5Al0.5-HT may be Mg6Fe1Al1(OH)16CO35H2O,
containing both Fe3+ and Al3+ as solid solutions, appeared which seems to be reasonable since the number of water
together with the weak line of Mg(Fe, Al)2O4 spinel. imp-Fe/ molecule is just between Mg6Fe2(OH)16CO36H2O [29] and
MgO also showed mainly the lines of periclase MgO Mg6Al2(OH)16CO34H2O [40].
The periclase Mg(Fe, Al)O phase was mainly detected
together with a small amount of spinel Mg(Fe, Al)2O4, in the
XRD patterns of Mg–Fe–Al samples after the calcination of
the precursors at 550 8C in both N2 and O2 atmospheres
(Fig. 1a and b). These results demonstrate that the mixed
oxide samples may mainly possess a brucite-like structure
with Fe3+ randomly distributed in the octahedral sites owing
to the ionic radius of Fe3+ (0.49 Å) being smaller than that of
Mg2+ (0.72 Å) [41]. We have reported that periclase
Mg(Ni2+, Al3+)O was formed by the calcination of Mg–
Ni–Al HT [25,26], where the ionic radii of Ni2+ and Al3+
were 0.69 and 0.39 Å, respectively [41]. It is possible to
consider the formation of Mg(Fe2+)O solid solutions, since
the ionic radius of Fe2+ is 0.78 Å [41]. Actually the
Mg(Fe2+)O solid solutions were formed by the reduction of
MgFe2O4 spinel during the preparation of Mg–Fe mixed
oxide from the HT precursors [42,43]. The formation of
Mg(Fe3+)O solid solutions was also reported in Mg–Fe
mixed oxide prepared from the HT precursors, by Shen and
co-workers [30] and Figueras and co-workers [33]. More-
over, the presence of Fe3+ in periclase Mg(Fe3+, Al3+)O can
be also supported by both Mössbauer spectra and X-ray
Fig. 1. XRD patterns of supported Fe catalysts effects of pre-treatment and absorption spectra in the present work (vide infra).
the reaction: (a) Mg3Fe0.5Al0.5 (after N2 treatment at 550 8C); (b)
Mg3Fe0.5Al0.5 (after O2 treatment at 550 8C); (c) Mg3Fe0.5Al0.5 (after H2 3.2. Electronic state of iron
treatment at 550 8C); (d) Mg3Fe0.5Al0.5 (after H2 treatment at 1000 8C); (e)
Mg3Fe0.5Al0.5 (a, after the reaction); (f) Mg3Fe0.5Al0.5 (d, after the reaction).
(&) Periclase Mg(Fe, Al)O; (&) Mg(Fe, Al)2O4 spinel; (+) Fe metal; (*) The color of the supported Fe catalysts is shown in
MgAl2O4; (*) Fe3C; (), graphite. Table 1 since it is a simple indication of whether bulk iron
224 Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231

oxide exists [44]. All as-synthesized Mg–Fe–Al HT samples


exhibited a white color, suggesting that no bulk iron oxide
existed and that all iron cations were incorporated inside the
framework of hydrotalcite after the co-precipitation. After
the calcination at 550 8C, both Mg3Fe0.5Al0.5 and
Mg2Fe0.25Al0.75 became off-white in color, suggesting the
presence of some extraframework iron. On the other hand,
the color from pale brown to brown of Mg2Fe0.75Al0.25,
Mg2Fe0.5Al0.5, and Mg2Fe with increasing color density
with increase in the Fe content suggests that these samples
contain aggregated iron oxide clusters; none the less, XRD
did not detect any iron oxide. For the two supported Fe
catalysts prepared by impregnation, imp-Fe/MgO was off-
white in color, suggesting the well-dispersed iron formation
on this catalyst, whereas imp-Fe/g-Al2O3 exhibited reddish
brown color probably due to the formation of bulk iron oxide
as detected by XRD.
As shown in the Mössbauer spectra of Mg3Fe0.5Al0.5
before and after the reaction and in the spectra of imp-Fe/
Fig. 2. Mössbauer spectra of supported Fe catalysts: (a) Mg3Fe0.5Al0.5
MgO (Fig. 2a–c), these samples displayed a doublet with an (before reaction); (b) Mg3Fe0.5Al0.5 (after reaction); (c) imp-Fe/MgO; (d)
isomer shift (IS) of around 0.3 mm s1 and with a imp-Fe/g-Al2O3.
quadrupole splitting (QS) of around 0.7 mm s1; such
values are typical of superparamagnetic Fe3+ species. The Al)2O4 spinel was observed as a minor phase or as an
spectrum of Mg3Fe0.5Al0.5 sample can be fitted with two amorphous phase (Fig. 1). It should be noted that the spinel
doublets, whose calculated parameters are reported in species was not strongly detected by XRD at the calcination
Table 2. One of the two components is attributed to the Fe3+ temperature of 550 8C. Moreover, the concentration of Fe3+
in Mg(Fe3+, Al)O as solid solutions, as seen in the XRD. will probably be higher in Mg(Fe3+, Al)2O4 spinel than in
Sometimes small ferric oxide particles or a-Fe2O3 were periclase Mg(Fe3+, Al)O, since Fe3+ possesses a regular site
observed in Mg–Fe [33] or Mg–Fe–Al [30] mixed oxides; in the former crystal structure but not in the latter. No clear
both showed the Fe3+ signal as a doublet in Mössbauer correlation was thus observed between the abundance of
spectra. Even though the off-white color suggests the Fe3+, shown as the value of relative contribution (Table 2),
possibility of iron oxide (Table 1), XRD exhibited no clear and the analytical results obtained by XRD. After the
evidence of such species and, moreover, a-Fe2O3 usually reaction, the Mg3Fe0.5Al0.5 sample exhibited a much broader
displays a sextet signal (vide infra). It is therefore most peak width and significantly larger QS in the Mössbauer
likely that the other doublet component observed in the spectrum. This is probably due to the fact that a part of Fe3+
Mössbauer spectra can be attributed to the Fe3+ in Mg(Fe3+, was reduced to Fe2+, as reported by Shen and co-workers
Al)2O4 spinel. The XRD patterns showed the presence of [30] and Figueras and co-workers [33]. The Mössbauer
periclase Mg(Fe3+, Al)O as a main phase, while Mg(Fe3+, spectrum of imp-Fe/MgO showed two doublets attributed to

Table 2
Mössbauer parameters of supported iron oxide catalysts
Catalyst Component Site ISa QSb LWc Bhfd Rel. contr. (%)e Phases by XRD
3+
Mg3Fe0.5Al0.5 Doublet Fe (1) 0.42 0.68 0.52 65.0 Mg(Fe, Al)2O4
Doublet Fe3+(2) 0.20 0.73 0.52 35.0 Mg(Fe, Al)O
Mg3Fe0.5Al0.5 (after reaction) Doublet Fe3+(1) 0.44 0.74 0.59 48.2 Mg(Fe, Al)2O4
Doublet Fe3+(2) 0.27 0.79 0.59 51.8 Mg(Fe, Al)O
imp-Fe/MgO Doublet Fe3+(3) 0.37 0.67 0.51 64.0 MgFe2O4
Doublet Fe3+(4) 0.15 0.77 0.51 36.0 Mg(Fe)O
imp-Fe/g-Al2O3 Doublet Fe3+(5) 0.51 0.88 0.67 18.9 Superpara Fe2O3
Doublet Fe3+(6) 0.26 0.89 0.67 63.7 Tetrahedral Fe-oxide
Sextet Fe3+(7) 0.36 0.22 0.50 50.1 17.4 a-Fe2O3
a
Isomer shift (mm s1).
b
Quadrupole splitting (mm s1).
c
Line width (mm s1).
d
Inner magnetic field (T).
e
Relative contribution.
Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231 225

both periclase Mg(Fe)O and MgFe2O4 spinel (Fig. 2c and


Table 2). It is supposed that the Fe3+ in Mg(Fe, Al)2O4 spinel
is more quickly reduced than that in periclase Mg(Fe, Al)O
during the reaction, since the Fe3+ ions in periclase Mg(Fe,
Al)O are much more highly diluted and probably better
stabilized. Actually, the reflection line of Mg(Fe, Al)2O4
spinel disappeared after the reaction (Fig. 1). It is thus
concluded that Fe3+(1) site is located in Mg(Fe, Al)2O4
spinel while Fe3+(2) site is in periclase Mg(Fe, Al)O
(Table 2). imp-Fe/g-Al2O3 showed the Mössbauer spectra
composed of two doublets and one sextet (Fig. 2d and
Table 2), while a-Fe2O3 (hematite) and g-Al2O3 alone were
observed in the XRD. The sextet of the site Fe3+(7) with IS
of 0.36 mm s1 is assigned to the Fe3+ in a-Fe2O3, as
reported in the study on Mg–Fe–Al mixed oxides; the sextet
has IS of 0.32 mm s1, characteristic of a-Fe2O3 [43]. One
of the other doublets, Fe3+(5), is assigned to the Fe3+ in
superpara Fe2O3 of fine particles, since the inner magnetic Fig. 4. Fe K-edge XANES spectra of supported Fe catalysts and materials
field of a-Fe2O3 was 50.1 T, a little smaller than the normal as control: (a) Fe2O3; (b) Mg3Fe0.5Al0.5-HT; (c) Mg3Fe0.5Al0.5; (d) imp-Fe/
values of 51–52 T. Another doublet, Fe3+(6), is assigned to MgO; (e) imp-Fe/g-Al2O3.
the Fe3+ in tetrahedral Fe oxide from the value of IS of
0.26 mm s1, since the IS of Fe3+O4 tetrahedra is generally samples prepared by CVD and assigned these bands to
observed at 0.2–0.32 mm s1 while that of Fe3+O6 octahedra isolated Fe3+ species interacting strongly with the support.
is generally observed at 0.35–0.55 mm s1. Mg3Fe0.5Al0.5 HT as prepared exhibited a peak at 270 nm
(Fig. 3b). This should probably be assigned to Fe3+ which is
3.3. Coordination sphere of iron isolated and octahedrally coordinated in brucite layered
structure [29]. Both Mg3Fe0.5Al0.5 and imp-Fe/MgO showed
Fig. 3 shows the diffuse reflectance UV–vis spectra. a- a broad peak around 330 nm together with a weak shoulder
Fe2O3 exhibited a peak at ca. 540 nm (Fig. 3a) assigned to around 480 nm. On imp-Fe/g-Al2O3, the latter shoulder
Fe2O3 particles. Centi and Vazzana [45] observed absorption shifted toward much longer wavelength and appeared at
bands at 270 and 330 nm in the UV–vis spectra of Fe/ZSM-5 540 nm. The peak around 330 nm is assigned to isolated
Fe3+ in either periclase Mg(Fe, Al)O or Mg(Fe, Al)2O4
spinel in Mg3Fe0.5Al0.5 mixed oxide. The isolated Fe3+ is
possibly formed also in imp-Fe/MgO, since neither XRD
reflection of Fe2O3 nor brown color was observed and,
moreover, no signal of Fe2O3 was detected in Mössbauer
spectra for this sample. The latter peaks around 480 nm for
Mg3Fe0.5Al0.5 and imp-Fe/MgO can probably be assigned to
the aggregated Fe3+ oxide clusters.
Fig. 4 shows the Fe K-edge XANES spectra of the
supported Fe catalysts, along with hematite (a-Fe2O3) as
control. The XANES spectra exhibit a pre-edge peak at
approximately 7107 eV, which is 10 eV below the midpoint
of the absorption band and is commonly assigned to iron in
tetrahedral coordination [46,47]. Iron atoms are principally
in octahedral coordination in both Mg3Fe0.5Al0.5 HT and a-
Fe2O3 and therefore showed the smallest pre-edge peak. If
one judges the relative intensities of the pre-edge peak for
Mg3Fe0.5Al0.5, imp-Fe/MgO, and imp-Fe/g-Al2O3 against
those for Mg3Fe0.5Al0.5 HT and a-Fe2O3, one will see that
the amount of iron in tetrahedral coordination must be rather
small and that a major part of iron exists in octahedral
coordination also in the former three catalysts. Fourier
Fig. 3. Uv–vis spectra of supported Fe catalysts: (a) Fe2O3; (b) transforms of the k3-weighted Fe K-edge EXAFS spectra of
Mg3Fe0.5Al0.5-HT; (c) Mg3Fe0.5Al0.5; (d) imp-Fe/MgO; (e) imp-Fe/g- the supported Fe catalysts along with hematite as control are
Al2O3. shown in Fig. 5. Mg3Fe0.5Al0.5 HT exhibited an intense peak
226 Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231

Fig. 6. Effect of Fe–Al composition on the dehydrogenation of ethylben-


zene over Mg2FexAl1x: (*) ethylbenzene conversion; (*) styrene selec-
tivity; (&) benzene selectivity; (~) toluene selectivity. Catalyst, 0.15 g; N2,
10 ml min1; ethylbenzene, 1.5 mmol h1; reaction temperature, 550 8C;
reaction time, 1 h.
Fig. 5. Fourier transforms of the k3-weighted Fe K-edge EXAFS spectra of
supported Fe catalysts and materials as control: (a) Fe2O3; (b) Mg3Fe0.5Al0.5-
HT; (c) Mg3Fe0.5Al0.5; (d) imp-Fe/MgO; (e) imp-Fe/g-Al2O3.
Fig. 6. In the absence of Fe, i.e., on Mg2Al mixed oxide, the
conversion of ethylbenzene was extremely low, and benzene
of Fe–O at 1.6 Å (non-phase shift corrected) together with a was produced as a by-product, suggesting that a thermal
weak peak at 2.6 Å (non-phase shift corrected) while a- b-fission reaction took place. Upon introducing Fe
Fe2O3 showed two strong peaks at 2.6 and 3.2 Å (non-phase component into the Mg2Al mixed oxide, we found that
shift corrected) in addition to that of the Fe–O bond. These the conversion of ethylbenzene increased and we observed
two peaks could be assigned the Fe–O–Fe linkages through selective formation of styrene. Increase in Fe content
edge-shared and corner-shared FeO6 octahedra [48]. resulted in an increase in the conversion of ethylbenzene, but
Mg3Fe0.5Al0.5 exhibited two peaks of Fe–O and Fe–O–Fe no substantial change was observed in the product
even though the intensities were significantly weakened selectivities. The activity of Mg3Fe0.5Al0.5 mixed oxide is
compared with those of Mg3Fe0.5Al0.5 HT, suggesting that shown in Table 3 together with those of imp-Fe/MgO and
the basic coordination structure around Fe was not changed imp-Fe/g-Al2O3. The selectivity to hydrogen was calculated
during the calcination. imp-Fe/MgO also showed almost based on the stoichiometry of the dehydrogenation of
identical spectra to those of Mg3Fe0.5Al0.5 with much lower ethylbenzene to styrene (1). The activity of Mg3Fe0.5Al0.5
intensities. On the other hand, imp-Fe/g-Al2O3 exhibited a
Ph-CH2 CH3 ! Ph-CH¼CH2 þ H2 (1)
peak at 3.2 Å (non-phase shift corrected) in addition to those
observed for Mg3Fe0.5Al0.5, among which the peaks mixed oxide was the highest among the catalysts tested
assignable to the Fe–O–Fe linkage were enhanced. The in the present work and was also higher than that obtained
latter result well coincided with the fact that the aggregated on Cr-MCM-41 in the presence of CO2 [49]. The selectivity
iron oxide clusters exist together with the isolated iron to hydrogen was high on Mg3Fe0.5Al0.5 mixed oxide and
species on imp-Fe/g-Al2O3 observed by both XRD and low on both imp-Fe/MgO and imp-Fe/g-Al2O3, none the
Mössbauer spectroscopies. less the selectivity to styrene was not affected on either
catalyst.
3.4. Activities of Mg–Fe–Al mixed oxides Gao and co-workers [35] reported that, in the dehy-
drogenation of ethylbenzene over Mg–Fe–Zn–Al mixed
The effect of Fe–Al composition in Mg2Fe1xAlx mixed oxide prepared from the corresponding HT, the catalytic
oxide in the dehydrogenation of ethylbenzene is shown in activity was enhanced in the presence of CO2 and afforded

Table 3
Ethylbenzene dehydrogenation over supported Fe catalystsa
Catalyst Conversion (%) of ethylbenzene Selectivity (%)
Styrene Benzene Toluene Hydrogenb
Mg3Fe0.5Al0.5 58.5 91.2 5.0 3.7 86.4
imp-Fe/g-Al2O3 34.0 92.8 4.6 2.6 70.0
imp-Fe/MgO 36.9 92.8 4.8 2.4 69.2
a
Reaction conditions: catalyst weight, 150 mg; reaction temperature, 550 8C; reaction time, 1 h; N2/Ar, 10/30 ml min1; ethylbenzene, 1.5 mmol h1.
b
Calculated based on the stoichiometry of ethylbenzene dehydrogenation to styrene.
Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231 227

Fig. 7. Effect of partial pressure of O2 on the dehydrogenation of ethyl-


benzene over Mg3Fe0.5Al0.5: (*) ethylbenzene conversion; (*) styrene Fig. 8. Effect of reaction temperature on the dehydrogenation of ethylben-
selectivity; (&) benzene selectivity; (~) toluene selectivity; (&) CO2 zene over Mg3Fe0.5Al0.5: (*) ethylbenzene conversion; (*) styrene selec-
selectivity; (~) CO selectivity. Catalyst, 0.15 g; pre-treated in N2 tivity; (&) benzene selectivity; (~) toluene selectivity. Catalyst, 0.15 g;
(20 ml min1) at 550 8C for 1 h; reaction in N2/Ar (10/30 ml min1) and pre-treated in N2 (20 ml min1) at 550 8C for 1 h; reaction in N2/Ar (10/
ethylbenezene (1.5 mmol h1) at 550 8C; reaction time, 1 h. 30 ml min1) and ethylbenezene (1.5 mmol h1); reaction time, 1 h.

the highest ethylbenzene conversion of 53.8% and a styrene Further increase in the reaction temperature to 600 8C
selectivity of 96.7% at 500 8C. Moreover, Mg–V–Al resulted in a decrease in the ethylbenzene conversion
mixed oxide prepared from the HT precursors showed high together with a drastic decrease in the styrene selectivity.
activity in the dehydrogenation of ethylbenzene in the Also at 600 8C, benzene formation was significantly
presence of CO2; the XRD analysis indicates the formation enhanced, suggesting an occurrence of thermal cracking
of Mg3V2O7 and Mg2Al2O4 in bulk, while the XPS results reactions of ethylbenzene. This was supported by the fact
point to the presence of a mixture of V5+ and V4+ ions on that significant coke formation was observed on the catalyst
the surface [27]. This Mg–V–Al mixed oxide showed after the reaction.
high activity also in the presence of O2; the highest
ethylbenzene conversion of 38% was obtained with a 3.5. Effect of pre-treatment on catalytic behavior of
styrene selectivity of 98% at 450 8C [28]. In the present Mg3Fe0.5Al0.5
work, addition of neither CO2 nor O2 showed any
favorable effect in the dehydrogenation of ethylbenzene Temperature programmed reduction (TPR) of
on Mg3Fe0.5Al0.5 mixed oxide. The effect of the addition Mg3Fe0.5Al0.5 mixed oxide showed two reduction peaks
of oxygen in the dehydrogenation of ethylbenzene on at 433 and 907 8C (Fig. 9). Shen and co-workers [43]
Mg3Fe0.5Al0.5 mixed oxide is shown in Fig. 7. With reported that the TPR of Mg2Fe1Al1 mixed oxide showed
increasing partial pressure of O2, the ethylbenzene two peaks at 506 and 904 8C. At the first TPR peak (506 8C),
conversion first decreased and then inversely increased hematite (a-Fe2O3) was reduced to Fe2+ and Fe0 ((2) and
accompanied by an increase in CO2 selectivity. Styrene (3)) while Mg(Fe, Al)2O4
formation was significantly suppressed in the presence of
O2, indicating that Mg3Fe0.5Al0.5 mixed oxide catalyzed a-Fe2 O3 þ MgO þ H2 ! Mg1x FeII O þ H2 O (2)
the combustion reaction and was not effective for the
oxydehydrogenation of ethylbenzene. The addition of a-Fe2 O3 þ H2 ! a-Fe þ H2 O (3)
CO2 simply resulted in a decrease in the ethylbenzene
conversion; none the less, the selectivities to styrene,
toluene, and benzene were not affected. Moreover, during
the reaction for 3 h, the ethylbenzene conversion gradually
decreased in the presence of CO2 while no substantial
decline was observed in the absence of CO2. These results
suggest that CO2 does not work as the oxidant on
Mg3Fe0.5Al0.5 mixed oxide as observed on Mg–V–Al
mixed oxide [28] or on Cr-MCM-41 [49].
The effect of the reaction temperature on the dehydro-
genation of ethylbenzene on Mg3Fe0.5Al0.5 mixed oxide is
shown in Fig. 8. When the reaction temperature was
increased from 500 to 575 8C, the ethylbenzene conversion Fig. 9. Temperature programmed reduction of Mg3Fe0.5Al0.5. Heating rate,
increased, while the styrene selectivity slowly decreased. 10 8C min1; in a mixture of 5 vol.% H2/Ar at a rate of 100 ml min1.
228 Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231

was transformed into Mg1xFexO (4). The second TPR peak and also that Mg3Fe0.5Al0.5 sample exhibited a much broader
(904 8C) corresponded to peak width and significantly larger QS in the Mössbauer
spectra after the dehydrogenation reaction (Table 2).
MgðFeIII ; AlÞ2 O4 þ H2 ! Mg1x FeIIx O þ H2 O þ MgðAlÞO After the pre-treatment in H2 atmosphere at 1000 8C, no
(4) reaction could be continued since the reactor was quickly
plugged by significant coke formation (Table 4). In the XRD
the reduction of Mg1xFexO4 to metallic Fe0 (5). The of the sample after the reduction (Fig. 1d), the lines of
reductions of Fe3+ to Fe2+ and Fe0 Mg(Al)O were sharpened while the lines of Fe metal
Mg1x FeIIx O þ H2 ! a-Fe þ H2 O þ MgO (5) appeared, indicating that periclase Mg(Fe, Al)O was
reductively decomposed to Mg(Al)O and Fe metal. When
were confirmed by both XRD and Mössbauer spectroscopy the reduced sample was used in the dehydrogenation
[42]. In the present work, the temperatures of both TPR reaction, the lines of MgO were further sharpened and, at the
peaks varied depending on the ratio of Mg/Fe/Al; the first same time, many reflection lines of Fe3C carbide appeared
peak significantly moved between 456 and 552 8C. Similar together with those of graphite and MgAl2O4 spinel in the
results were obtained for Mg–Fe mixed oxides prepared XRD (Fig. 1f). These results indicate that periclase Mg(Al)O
from HT precursors [29]. It is therefore possible that Fe3+ in was decomposed into MgO and MgAl2O4 spinel and,
Mg3Fe0.5Al0.5 mixed oxide was reduced to Fe2+ or Fe0 by H2 moreover, that Fe metal was converted into Fe3C carbide by
formed during the dehydrogenation reaction at 550 8C. coke deposited on it during the reaction.
Effects of the conditions of pre-treatment of Mg3Fe0.5Al0.5
mixed oxide on the activity in the dehydrogenation reaction 3.6. Possible active phase on Mg3Fe0.5Al0.5
are shown in Table 4. The pre-treatment in H2 atmosphere was
carried out under the same conditions as that of TPR to exactly The time course of the dehydrogenation reaction is shown
follow the effect of the reduction of Fe3+ to Fe2+ or Fe0. in Fig. 10. After the pre-treatment of Mg3Fe0.5Al0.5 mixed
Sometimes the reduction treatment with H2 brought over- oxide in N2 atmosphere (Fig. 10, full line), a short induction
reduction to the catalyst, resulting in the formation of metallic period of 20 min was observed, i.e., the conversion of
Fe. As a result, coking became significant on the metallic Fe ethylbenzene started at a low value and increased during
and the reaction was finally suppressed (vide infra). In the 20 min to reach a steady state value. The selectivity to
XRD of the samples, reflection lines of periclase Mg(Fe, Al)O styrene gradually increased while that to benzene decreased
were observed in all samples treated at 550 8C (Fig. 1a–c). A during the reaction for 3 h. When the mixed oxide was pre-
weak reflection of Mg(Fe, Al)2O4 spinel was observed for the treated in H2 atmosphere, the induction period almost
samples treated in both N2 and O2 atmospheres (Fig. 1a and b), disappeared and both conversion and selectivity showed
whereas no such reflection was observed for the sample high values, close to those of the steady state, even at the
treated in H2 atmosphere (Fig. 1c). Even with such changes in beginning of the reaction (Fig. 10, dotted line). After the
the XRD, no significant effect of pre-treatments in N2, H2, and reaction reached steady state, the reaction profiles were
O2 atmosphere at 550 8C was observed in the reaction results. almost the same as those obtained after the pre-treatment in
The pre-treatments in both N2 and H2 atmosphere at 550 8C N2 atmosphere. It was thus clearly confirmed that the
resulted in a similar high activity, whereas that in O2 reduction pre-treatment brought a favorable effect to the
atmosphere at 550 8C caused a little decrease in the activity. catalytic activity of Mg3Fe0.5Al0.5 catalyst.
These results suggest that the surface of Mg3Fe0.5Al0.5 mixed Many papers suggest that the active phase in iron oxide
oxide was almost same after 1 h of the dehydrogenation catalyst is Fe3+. Ranke co-workers [14] reported that
reaction, i.e., the surface was reduced by H2 produced during unpromoted Fe2O3 was much more active than Fe3O4 and
the reaction. This was actually seen in the facts that the that KFeO2 is the active species for K-promoted iron oxide
reflection line of Mg(Fe, Al)2O4 spinel disappeared (Fig. 1e) catalyst in the dehydrogenation of ethylbenzene. KFeO2 as

Table 4
Effect of pre-treatment of Mg3Fe0.5Al0.5 mixed oxide catalyst
Condition of pre-treatment Conversion (%) of ethylbenzene Selectivity (%)
Styrene Benzene Toluene Hydrogena
N2b 60.0 91.7 4.7 3.6 92.0
O2c 58.5 91.2 5.0 3.7 86.4
H2d 59.1 93.2 4.0 2.8 92.0
H2e – – – – –
a
Calculated based on the stoichiometry of ethylbenzene dehydrogenation to styrene.
b
With N2 (20 ml min1) at 550 8C for 1 h.
c
With O2/N2 (10/10 ml min1) at 550 8C for 1 h.
d
With H2/Ar (5/95 ml min1) from room temperature to 550 8C at a rate of 10 8C min1.
e
With H2/Ar (5/95 ml min1) from room temperature to 1000 8C at a rate of 10 8C min1.
Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231 229

Fig. 10. Time course of the dehydrogenation of ethylbenzene over


Mg3Fe0.5Al0.5: (*) ethylbenzene conversion; (*) styrene selectivity; (&)
benzene selectivity; (~) toluene selectivity. Solid line, pre-treated in N2
(20 ml min1) at 550 8C for 1 h; dotted line, pre-treated in H2/Ar (5/
95 ml min1) from room temperature to 550 8C at 10 8C min1. Catalyst, Fig. 11. Fe K-edge XANES spectra of Mg3Fe0.5Al0.5 after the reaction: (*)
0.15 g; reactionin N2/Ar (10/30 ml min1) and ethylbenezene (1.5 mmol h1) Mg3Fe0.5Al0.5 after the reaction for 1 h; (~) Mg3Fe0.5Al0.5 before the
at 550 8C. reaction; (&) Fe(II)O.

the active phase was also identified using SEM and Auger gave clear evidence of the fact that defective surfaces are
microscopy by Lundin et al. [7]. Moreover, a study on K- more active than well-ordered ones [13]. Also, the in situ
promoted iron oxide catalyst films prepared on Ru(0 0 0 1) study of MCM-41-supported iron oxide by XANES and
as a model catalyst reported that the KFeO2 shell around a EXAFS showed that a distorted form of iron oxide species is
K2Fe22O34 core was identified for the active phase [50]. In metastable and contains labile surface oxide anions, which
the ethylbenzene dehydrogenation over K-promoted iron are probably responsible for the high initial catalytic activity
oxide catalyst, the fully oxidized iron phases containing Fe3+ during ethylbenzene dehydrogenation reaction at 500 8C
ions are responsible for high catalytic activity; their partial [56]. Such effects of reduced iron species or defect structure
reductions by generated H2 caused a deactivation, and the due to the partial reduction on the catalyst surface are likely
treatment with steam leads to a partial re-oxidation, resulting in the present work.
in the re-activation of the catalyst [51]. It was also reported
that the activity losses over the Fe-Mg-Al catalysts with 3.7. Possible roles of Mg and Al
time-on-stream were completely restored by oxygen pulses
[34]. The addition of Mg as well as Al will be also important in
Fe K-edge XANES spectra of Mg3Fe0.5Al0.5 after the the high activity of Mg–Fe–Al mixed oxides. A high
reaction, along with that before the reaction and that of stability was observed by the addition of Mg on K-promoted
Fe(II)O as control, are shown in Fig. 11. The iron species iron oxide catalyst; it was attributed to the fact that MgO
was slightly reduced after the reaction; the charge of iron is prevented the pore destruction of the catalyst [19]. By the
considered to be between Fe2+ and Fe3+ judging from the process of supporting promoted iron oxide catalyst on MgO,
location of the absorbance curve of Mg3Fe0.5Al0.5 after the coke formation was limited because the catalyst was
reaction in the spectra. In the present work, it must be subjected to a suitable pre-treatment [20]. Aluminum was
emphasized that the reduced iron species is active as well as proved to be an excellent promoter, preventing sintering in
stable in the ethylbenzene dehydrogenation (Figs. 10 and iron-oxide catalysts [17,18]. The supported system can
11). It is noteworthy that H2 is formed during the prevent the physical degradation of the catalyst particles.
dehydrogenation and reduces the surface of catalyst [52]. In the present work, the surface area of Mg3Fe0.5Al0.5
Reports also suggest the possibility of Fe2+ as the active was 169.0 m2 g1 cat before the reaction, and it decreased
species. Lee [53] observed that a-Fe2O3 (hematite) as a main to 152.4 and 147.1 m2 g1 cat after the reaction for 1 and 3 h,
component in the catalyst is reduced to Fe3O4 (magnetite) respectively, suggesting that the surface area was somewhat
during the reaction and that the latter is more selective. stabilized during the reaction.
Courty and Le Page [54] mentioned that the catalyst is more The pore-size of Mg3Fe0.5Al0.5 showed a wide distribu-
stable after a reduction of Fe2O3 to Fe3O4 after the first 200 h tion from 2 nm to above 25 nm as well as an increasing
operation. Yang et al. [55], using Mössbauer spectroscopy distribution with an increase in radius (Rp). The lowest limit
on K-promoted iron oxide catalyst, suggested that the active of pore radius detectable in the present method is 2 nm, and
site contained no K and was promoted by a rapid electron therefore no information concerning micropores smaller
exchange between Fe2+ and Fe3+ catalyzed by K. Use of than 2 nm could be obtained. However, no significant
unpromoted Fe2O3 in the ethylbenzene dehydrogenation distribution less than 5 nm was observed on Mg3Fe0.5Al0.5.
230 Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231

Mg3Fe0.5Al0.5 calcined at 850 8C showed a sharp distribu- processes. The high catalytic performance is due to the
tion at less than 5 nm, together with a broad distribution formation of partially reduced iron oxides on the surface of
around 20 nm in radius. When the calcination temperature catalyst, together with the high surface area along with the
was increased, the pores with large size diminished and the porosity. All these characters are owing to the fact that
pore-size distribution tended to converge to the small values. catalysts were prepared via periclase Mg(Al, Fe)O as solid
Mg–Al(3/1) hydrotalcite showed a peak around 2–3 nm solutions intermediate starting from Mg–Fe–Al hydrotalcite
together with a wide distribution up to 20 nm in radius. The as the precursors.
former pore-sizes are probably related to the layered
structure, while the latter is due to a ‘‘card house’’ structure
consisting of many small plates. We calculated the basal References
interlayer spacing from the strong symmetric (0 0 3)
reflection (2u = 13.48) of Mg–Al(3/1) hydrotalcite. If the [1] E.H. Lee, Catal. Rev. Eng. Sci. 8 (1973) 285.
thickness of the brucite-like layer is assumed to be 4.8 Å [2] K. Kochloefl, in: G. Ertl, H. Knözinger, J. Weitkamp (Eds.),
Handbook of Heterogeneous Catalysis, vol. 5, VCH, Weinheim,
[57], the interlayer distance corresponds to 2.9 Å. The high 1998, p. 2151.
surface area of Mg3Fe0.5Al0.5 is probably due to the presence [3] T. Hirano, Appl. Catal. 26 (1986) 65.
of meso- and macro-pores in the catalyst. The pore-size [4] T. Hirano, Appl. Catal. 28 (1986) 119.
distribution of Mg3Fe0.5Al0.5 showed no substantial change [5] M. Muhler, J. Schütze, M. Wesemann, T. Rayment, A. Dent, R.
from before to after the reaction. Schlögl, G. Ertl, J. Catal. 126 (1990) 339.
[6] M. Muhler, R. Schlögl, G. Ertl, J. Catal. 138 (1992) 413.
The amounts of coke deposited on the catalyst were 15.7 [7] J. Lundin, L. Holmlid, P.G. Menon, L. Nyberg, Ind. Eng. Chem. Res.
and 22.5 mg g1 cat after 1 h of reaction in CO2 and Ar, 32 (1993) 2500.
respectively; these amounts were far larger than the values, [8] K. Coulter, D.W. Goodman, R.G. Moore, Catal. Lett. 31 (1995) 1.
i.e., 1.83 and 0.60 mg g1 cat , observed after 5 h of reaction
[9] S.C. Ndlela, B.H. Shanks, Ind. Eng. Chem. Res. 42 (2003) 2112.
[10] W.D. Mross, Catal. Rev.-Sci. Eng. 24 (1983) 591.
over Cr-MCM-41-DHT catalysts in CO2 and He, respec-
[11] R.L. Hirsch, J.E. Gallagher, R.R. Lessard, R.D. Wesselhoft, Science
tively [58]. These values suggest that the coking was not 215 (1982) 121.
main reason of the catalyst deactivation, since no deactiva- [12] W.P. Addiego, C.A. Estrada, D.W. Goodman, M.P. Rosynek, R.G.
tion was observed over Mg3Fe0.5Al0.5 catalyst while a clear Windham, J. Catal. 146 (1994) 407.
deactivation took place over Cr-MCM-41-DHT catalysts [13] W. Weiss, D. Zscherpel, R. Schlögl, Catal. Lett. 52 (1998) 215.
[58], and, moreover, the deactivation of Mg3Fe0.5Al0.5 [14] O. Shekhah, W. Ranke, R. Schlögl, J. Catal. 225 (2004) 56.
[15] B.D. Herzog, H.F. Raso, Ind. Eng. Chem. Prod. Res. Dev. 23 (1984)
catalyst occurred in CO2 but was not observed in Ar 187.
atmosphere. The porous structure with the high surface area [16] A.B. Styles, in: B.E. Leach (Ed.), Applied Industrial Catalysis, Aca-
was sufficiently preserved and no remarkable deactivation demic Press, New York, 1983, p. 137.
was observed on Mg3Fe0.5Al0.5 during the reaction for 3 h. [17] G.C. Araújo, M.C. Rangel, Catal. Today 62 (2000) 201.
These stabilizing effects on the catalytic activity are [18] H. Topsoe, J.A. Dumesic, M. Boudart, J. Catal. 28 (1978) 477.
[19] T. Hirano, Bull. Chem. Soc. Jpn. 59 (1986) 2672.
probably due to the fact that catalyst was prepared via [20] D.E. Stobbe, F.R. van Buren, A.W. Stobbe-Kreemers, A.J. van Dillen,
solid solutions, i.e., periclase Mg(Fe, Al)O, starting from J.W. Geus, J. Chem. Soc., Faraday Trans. 87 (1991) 1631.
Mg–Fe–Al hydrotalcite as the precursors. [21] A.C. Oliveira, A. Valentini, P.S.S. Nobre, M.C. Rangel, React. Kinet.
Catal. Lett. 75 (2002) 135.
[22] F. Cavani, F. Trifiro, A. Vaccari, Catal. Today 11 (1991) 173.
[23] O. Clause, M. Gazzano, F. Trifiro, A. Vaccari, L. Zatorski, Appl. Catal.
4. Conclusion 73 (1991) 217.
[24] E.C. Kruissink, L.L. Van Reijen, J.R.H. Ross, J. Chem. Soc., Faraday
Mg-Fe-Al mixed oxide catalysts prepared from the Trans. 177 (1991) 649.
corresponding layered double hydroxides as the precursors [25] K. Takehira, T. Shishido, P. Wang, T. Kosaka, K. Takaki, Phys. Chem.
were successfully applied for the ethylbenzene dehydro- Chem. Phys. 5 (2003) 3801.
[26] K. Takehira, T. Shishido, P. Wang, T. Kosaka, K. Takaki, J. Catal. 221
genation to styrene. By calcining the precursors, we (2004) 43.
obtained the mixed oxides with a high surface area and a [27] G. Carja, R. Nakamura, T. Aida, H. Niiyama, J. Catal. 218 (2003)
mesoporous character, where periclase Mg(Fe, Al)O was 104.
mainly detected by the XRD analysis. During the reaction, [28] F. Malherbe, C. Forano, B. Sharma, M.P. Atkins, J.P. Besse, Appl. Clay
Sci. 13 (1998) 381.
the iron species on the catalyst surface was partially reduced
[29] J. Shen, B. Guang, M. Tu, Y. Chen, Catal. Today 30 (1996) 77.
to the valence state between Fe2+ and Fe3+, and the partially [30] M. Tu, J. Shen, Y. Chen, J. Solid State Chem. 128 (1997) 73.
reduced iron species were effective for the dehydrogenation. [31] P.S. Kumbhar, J. Sanchez-Valente, F. Figueras, Tetrahedron Lett. 39
Neither CO2 nor O2 showed any favorable effect as the (1998) 2573.
additive in the reaction medium in the dehydrogenation [32] S.M. Auer, J.-D. Grunwaldt, R.A. Köppel, A. Baiker, J. Mol. Catal. A
reaction. The ethylbenzene conversion of 60% and the 139 (1999) 305.
[33] P.S. Kumbhar, J. Sanchez-Valente, J.M.M. Millet, F. Figueras, J. Catal.
styrene selectivity of 95% were stably observed over 191 (2000) 467.
Mg3Fe0.5Al0.5 catalyst at the optimum temperature of [34] P. Kuśtrowski, A. Rafalska-Łasocha, D. Majda, D. Tomaszewska, R.
550 8C, which was lower than that employed in commercial Dziembaj, Solid State Ionics 141–142 (2001) 237.
Y. Ohishi et al. / Applied Catalysis A: General 288 (2005) 220–231 231

[35] X. Ye, N. Ma, W. Hua, Y. Yue, C. Miao, Z. Xie, Z. Gao, J. Mol. Catal. [47] S.A. Axon, K.K. Fox, S.W. Carr, J. Klinowski, Chem. Phys. Lett. 189
A 217 (2004) 103. (1992) 1.
[36] N. Mimura, I. Takahara, M. Saito, Y. Sasaki, K. Murata, Catal. Lett. 78 [48] Y. Ma, S.L. Suib, Chem. Mater. 11 (1999) 3545.
(2002) 125. [49] Y. Ohishi, T. Kawabata, T. Shishido, K. Takaki, Q. Zhang, Y. Wang, K.
[37] M. Miyata, A. Okada, Clays Clay Miner. 25 (1977) 14. Takehira, J. Mol. Catal. A 230 (2005) 49.
[38] K. Tanaka, H. Yamashita, R. Tsuchitani, T. Funabiki, T. Yoshida, J. [50] G. Ketteler, W. Ranke, R. Schlögl, J. Catal. 212 (2002) 104.
Chem. Soc., Faraday Trans. 84 (1988) 2987. [51] X.M. Zhu, M. Shön, U. Bartmann, A.C. van Veen, M. Muhler, Appl.
[39] D. Dollimore, G.R. Heal, J. Appl. Chem. 14 (1964) 109. Catal. A 266 (2004) 99.
[40] A. Drits, T.N. Sokolova, G.V. Sokolova, V.I. Cherkashin, Clays Clay [52] G.R. Meima, P.G. Menon, Appl. Catal. A 212 (2001) 239.
Miner. 35 (1987) 401. [53] E.H. Lee, Catal. Rev. 8 (1973) 285.
[41] R.D. Shannon, Acta Crystallogr. A 32 (1976) 751. [54] Ph. Courty, J.F. Le Page, Stud. Surf. Sci. Catal. 3 (1979) 293.
[42] L.A. Boot, A.J. van Dillen, J.W. Geus, A.M. van der Kraan, A.A. vand [55] X. Yang, S. Weng, K. Jiang, L. Mao, Y. Euong, K. Jing, Hyperfine
der Horst, F.R. van Buren, Appl. Catal. A 145 (1996) 389. Interact. 69 (1991) 863.
[43] X. Ge, M. Li, J. Shen, J. Solid State Chem. 161 (2001) 38. [56] S.-T. Wong, J.-F. Lee, S. Cheng, C.-Y. Mou, Appl. Catal. A 198 (2000)
[44] P. Ratnasami, R. Kumar, Catal. Today 9 (1991) 329. 115.
[45] G. Centi, F. Vazzana, Catal. Today 53 (1999) 683. [57] M.A. Drezdon, Inorg. Chem. 37 (1988) 4628.
[46] D.W. Lewis, G. Sanker, C.R.A. Catlow, S.W. Carr, J.M. Thomas, Nucl. [58] K. Takehira, Y. Ohishi, T. Shishido, T. Kawabata, K. Takaki, Q. Zhang,
Instrum. Methods Phys. Res. B 97 (1995) 44. Y. Wang, J. Catal. 224 (2004) 404.

Вам также может понравиться